• No results found

Insights into thermoadaptation and the evolution of mesophily from the bacterial phylum Thermotogae

N/A
N/A
Protected

Academic year: 2022

Share "Insights into thermoadaptation and the evolution of mesophily from the bacterial phylum Thermotogae"

Copied!
41
0
0

Laster.... (Se fulltekst nå)

Fulltekst

(1)

1 Insights into the evolution of mesophily from the bacterial phylum Thermotogae

1 2

Stephen M. J. Pollo, Olga Zhaxybayeva, Camilla L. Nesbø 3

4

Stephen M. J. Pollo 5

Department of Biological Sciences, 11455 Saskatchewan Drive, University of Alberta, Edmonton, 6

Alberta, Canada, T6G 2E9. pollo@ualberta.ca 7

Olga Zhaxybayeva 8

Department of Biological Sciences and Department of Computer Science, Dartmouth College, 78 9

College Street, Hanover, NH, 03755, U.S.A. olga.zhaxybayeva@dartmouth.edu 10

Camilla L. Nesbø 11

Department of Biological Sciences, 11455 Saskatchewan Drive, University of Alberta, Edmonton, 12

Alberta, Canada, T6G 2E9 and Centre for Ecological and Evolutionary Synthesis (CEES), 13

Department of Biology, University of Oslo, P.O. Box 1066 Blindern, N-0316 Oslo, 14

Norway. nesbo@ualberta.ca and c.l.nesbo@bio.uio.no 15

16

Corresponding Author:

17

Camilla L. Nesbø 18

Department of Biological Sciences, CW 405 Biological Sciences Bldg., 11455 Saskatchewan 19

Drive , University of Alberta, Edmonton, Alberta, Canada, T6G 2E9. Telephone: (01) 780-492- 20

8956 nesbo@ualberta.ca and c.l.nesbo@bio.uio.no 21

Running Title: Temperature adaptation in Thermotogae 22

(2)

2 Abstract: Thermophiles are extremophiles that grow optimally at temperatures > 45°C. In order 23

to survive and maintain function of their biological molecules, they have a suite of characteristics 24

not found in organisms that grow at moderate temperature (mesophiles). At the cellular level, 25

thermophiles have mechanisms for maintaining their membranes, nucleic acids and other cellular 26

structures. At the protein level, each of their proteins remains stable and retains activity at 27

temperatures that would denature their mesophilic homologs. Conversely, cellular structures and 28

proteins from thermophiles may not function optimally at moderate temperatures. These 29

differences between thermophiles and mesophiles presumably present a barrier to evolutionary 30

transitioning between the two lifestyles. Therefore, studying closely related thermophiles and 31

mesophiles can help us determine how such lifestyle transitions may happen. The bacterial 32

phylum Thermotogae contains hyperthermophiles, thermophiles, mesophiles and organisms with 33

temperature ranges wide enough to span both thermophilic and mesophilic temperatures.

34

Genomic, proteomic and physiological differences noted between other bacterial thermophiles 35

and mesophiles are evident within the Thermotogae. We argue that the Thermotogae is an ideal 36

group of organisms for understanding both the response to fluctuating temperature as well as 37

long-term evolutionary adaptation to a different growth temperature range.

38 39

Key Words: lateral gene transfer, Kosmotoga, Mesotoga, thermostability, stress response 40

41

(3)

3 Introduction

42

Extremophiles are organisms that thrive under extreme environmental conditions unsuitable for 43

survival of most other organisms. As such, they are of great interest for delineating the limits of 44

conditions that permit life’s existence, a key insight needed to advance efforts in the search for 45

life on Earth and other planets (Pikuta et al. 2007; Rothschild and Mancinelli 2001). Additionally, 46

due to their intrinsically “extreme” nature, these organisms are also desirable sources of enzymes 47

and other biomolecules that function under conditions that render other organisms and their 48

enzymes inactive. Such biomolecules may have a wide range of biotechnological and industrial 49

applications from clean energy to bioremediation and carbon sequestration.

50

When examining temperature as a parameter that can either permit or exclude life, there 51

are mesophiles, the organisms that grow optimally at moderate temperatures, and two types of 52

extremophiles: psychrophiles, which grow optimally at temperatures below 15°C, and 53

thermophiles, which grow optimally at temperatures above 45°C (Kimura et al. 2013). Within 54

thermophiles, organisms growing optimally at > 80°C are commonly referred to as 55

hyperthermophiles. Thermophiles are of particular interest due to their ability to withstand the 56

denaturing effect of higher temperatures on biological molecules such as proteins and DNA (Li et 57

al. 2005).

58

The phylogenetic position of the hyperthermophile-containing bacterial lineages 59

Thermotogae, Thermodesulfobacteria and Aquificae at, or close to, the base of the 16S rRNA tree 60

of life (Fig. 1), has been used as support for the hypothesis that the ancestor of the bacterial 61

domain was a hyperthermophile (Achenbach-Richter et al. 1987). Similarly, thermophilic 62

Archaea are also found at the base of the Archaeal domain (Fig. 1). Together with the proposed 63

high temperature conditions of early Earth this led to the hypothesis that the last universal 64

(4)

4 common ancestor (LUCA) was a hyperthermophile (Pace 1991). A (hyper)thermophilic LUCA is 65

also supported by experimental evidence from resurrection of ancestral nucleoside diphosphate 66

kinases and characterizing their properties (Akanuma et al. 2013). Other lines of evidence, 67

however, suggest that the LUCA may have been either a mesophile or a thermophile growing 68

optimally below 80°C (Boussau et al. 2008; Brochier-Armanet and Forterre 2006). Whether the 69

LUCA lived at the time of life's origin or much later remains debatable as well (Zhaxybayeva and 70

Gogarten 2004).

71

Regardless of the optimal growth temperature of the LUCA, the ancestors of present day 72

bacterial and archaeal lineages have had to modify their cellular structures and protein 73

compositions to transition between mesophilic and thermophilic lifestyles (Boussau et al. 2008).

74

Given the distribution of mesophiles and thermophiles on the Tree of Life (Fig. 1), we infer that 75

such transitions likely happened independently multiple times. This same inference has been 76

made based on multivariate analyses of the amino acid compositions of 279 prokaryotes (Puigbò 77

et al. 2008) and from the different mechanisms of DNA supercoiling and the phylogeny of the 78

involved genes (López-García 1999). This conjecture is also supported by reconstruction and 79

synthesis of ancestral versions of enzymes and examining the optimal temperature at which they 80

function. For example, examination of LeuB enzymes (3-isopropylmalate dehydrogenase) in the 81

Bacillus genus suggests multiple transitions between thermophilic and mesophilic temperature 82

optima when going forward in evolutionary time from the Bacillus ancestor (Hobbs et al. 2012).

83

Therefore, thermophily has been lost and gained throughout the evolutionary history of the genus 84

Bacillus. Similarly, analysis of extant and reconstructed ancestral myo-inositol-3-phosphate 85

synthase enzymes from Thermotoga and Thermococcales suggests higher optimal growth 86

temperatures of the ancestors (Butzin et al. 2013), indicating fluctuations of the tolerated 87

(5)

5 temperature ranges of these organisms throughout their evolutionary history. Together these 88

studies imply that temperature adaptations may not be too difficult, and the growth temperature 89

range may change rapidly and frequently in many lineages.

90

Temperature adaptation can be defined either as a response of an individual cell to 91

changes in temperature, or as an evolutionary adaptation of an organismal lineage (such as 92

‘species’) to growth within a certain temperature range. To distinguish between the two, we will 93

refer to temperature response for the former and temperature adaptation for the latter. These two 94

phenomena are related, as selection acting on temperature responses may eventually lead to 95

temperature adaptations. In this review we focus on organismal responses and lineage adaptations 96

to moderate and high temperatures. For a review of adaptation to very low growth temperatures 97

see Siddiqui et al. (2013). Specifically, we will discuss properties of thermophiles, and how these 98

properties may relate to a transition between thermophily and mesophily, with a particular 99

emphasis on the bacterial phylum Thermotogae.

100 101

The Thermotogae 102

Bacteria belonging to the Thermotogae phylum were first isolated by Karl Stetter and colleagues 103

in 1986 from geothermally heated sea floors (Huber et al. 1986). Their name derives from the 104

unique outer sheath-like structure that balloons over each end of the cell, known as the “toga”

105

(Fig. 2) (Huber et al. 1986). There are 12 described genera in this phylum, most of which are 106

thermophiles (Fig. 3). In the accepted taxonomy, these genera are all grouped in a single order, 107

Thermotogales, and one family, Thermotogaceae. However, a reclassification of these bacteria 108

into separate orders is overdue, and a division into three orders and four families has been 109

(6)

6 recently proposed (Bhandari and Gupta (2014); Fig. 3). While the new classification is based on 110

conserved indels, it is consistent with the 16S rRNA phylogeny (Fig. 3).

111

Thermotogae are anaerobes and organotrophs, capable of growing on a wide range of 112

complex substrates (Conners et al. 2006). They are found in hot ecosystems all over the world 113

including thermal springs, hydrothermal vents, and petroleum reservoirs (Huber and Hannig 114

2006; Ollivier and Cayol 2005), with some members growing at temperatures up to 90°C.

115

Although it was long thought that the Thermotogae only harbored thermophiles and 116

hyperthermophiles (11 of 12 genera are entirely composed of thermophiles or 117

hyperthermophiles) (Fig. 3), mesophilic Thermotogae from the genus Mesotoga have recently 118

been detected and isolated from cool hydrocarbon-impacted sites such as oil reservoirs and 119

polluted sediments (Ben Hania et al. 2011; Ben Hania et al. 2013; Nesbø et al. 2006b; Nesbø et 120

al. 2010; Nesbø et al. 2012). Interestingly, the closest relative of Mesotoga, Kosmotoga olearia, 121

has an unusually wide growth temperature range, which may have been important in Mesotoga’s 122

adaptation to low temperature (DiPippo et al. 2009; Nesbø et al. 2012).

123

As of May 2015, over 80 completed and ongoing Thermotogae genome projects 124

comprise 10 of the 12 described Thermotogae genera, with no genome projects for Geotoga nor 125

Oceanotoga (Benson et al. 2014; Reddy et al. 2014). The maximum divergence in the 16S rRNA 126

genes of these cultivated Thermotogae is ~25%, similar to what is observed for other bacterial 127

phyla (Konstantinidis and Tiedje 2005). For protein coding genes pairwise average amino acid 128

identity (AAI; Konstantinidis and Tiedje 2005) between genera ranges from 45 to 69% (average 129

49%). Phylogenetic analysis of environmental 16S rRNA gene sequences shows several novel 130

Thermotogae lineages without any cultivated members, and based on the nucleotide identity they 131

would be classified as new genera (Nesbø et al. 2010). Thus, as with most microbial lineages, 132

(7)

7 there is a large unknown diversity of Thermotogae. At least four of these new lineages have only 133

been detected in low temperature environments (as low as 9.5°C), suggesting that Thermotogae 134

might be common in mesothermic environments. Interestingly, on the phylogenetic tree these 135

likely mesophilic lineages fall within multiple thermophilic clades (Nesbø et al. 2010), 136

suggesting several independent adaptations to lower temperatures.

137

With mesophilic Thermotogae only recently discovered, the functional characterization of 138

this phylum has focused on thermophiles, mainly the hyperthermophilic organisms Thermotoga 139

maritima and Thermotoga neapolitana. Protein crystal structures have also been experimentally 140

determined for a large portion of the T. maritima proteome (DiDonato et al. 2004; Lesley et al.

141

2002), and the protein structures of its central metabolic networks were modeled by Zhang et al.

142

(2009). Complimented with models of high temperature hydrogen and sulfur metabolism 143

(Cappelletti et al. 2014; Schut et al. 2012), this wealth of functional information makes the 144

Thermotogae a promising microbial lineage for industrial and biotechnological applications. For 145

example, most Thermotogae produce hydrogen that may be harvested (e.g., Nguyen et al. (2008) 146

and Maru et al. (2012)). The hydrogen production of T. maritima can be boosted via metabolic 147

engineering, as was demonstrated by an in silico re-design of its metabolism (Nogales et al. 2012).

148

Additionally, while the degradation of sugars by many Thermotogae results in the production of 149

CO2 and acetate, T. neapolitana has been shown to convert these by-products to lactic acid when 150

grown in a CO2 atmosphere, a process suggested to have potential in carbon capture (D'Ippolito et 151

al. 2014).

152

Carbohydrate utilization by T. maritima has been examined by studying the substrate 153

specificities and affinities of its carbohydrate transporters (Boucher and Noll 2011; Cuneo et al.

154

2009; Ghimire-Rijal et al. 2014; Nanavati et al. 2005; Nanavati et al. 2006) and their 155

(8)

8 transcriptional regulation in response to growth on different saccharides (Frock et al. 2012).

156

Information about substrate specificities, enzymatic activities and catalytic mechanisms of many 157

of T. maritima’s glycoside hydrolases are also available (Arti et al. 2012; Comfort et al. 2007;

158

Kleine and Liebl 2006), which has been used, for instance, to engineer an alpha-galactosidase 159

from T. maritima into an efficient alpha-galactosynthase (Cobucci-Ponzano et al. 2011). The 160

transcriptional regulation of glycoside hydrolases and other carbohydrate metabolism-related 161

genes in response to growth on various carbohydrates highlights the differences in carbohydrate 162

utilization, even between closely related Thermotogae lineages (Chhabra et al. 2002; Chhabra et 163

al. 2003; Frock et al. 2012). Moreover, interconnections exist between sugar regulons in T.

164

maritima’s carbohydrate utilization network, suggestingcoordinated regulatory responses to 165

particular types of complex carbohydrates (Rodionov et al. 2013). This rich knowledge base will 166

be very useful in comparative studies of thermophilic and mesophilic Thermotogae lineages and, 167

ultimately, will lead to understanding processes leading to shifts in an organism’s growth 168

temperature range.

169 170

General cellular adaptations to thermophily 171

Regardless of whether cells are responding to transient temperature increases within their growth 172

range or evolving to an alternate growth range, changes in temperature require major 173

modifications across the cell to optimize cell function and growth. The following sections discuss 174

some of these temperature responses and adaptations in microbial cells.

175 176

The effect of temperature on cellular membranes: maintaining a fluid envelope 177

(9)

9 The cell membrane is critical to cell function since it maintains and separates the interior cell 178

environment from the exterior environment. In order to serve its function, a lipid membrane must 179

be impermeable to most solutes and maintain a liquid crystalline phase, even under stress (de 180

Mendoza 2014). As the temperature increases, membrane integrity and impermeability become 181

compromised, which eventually results in cell death (Chang 1994). Therefore, thermophiles must 182

maintain their membranes under conditions that could inactivate those of mesophiles. Bacteria 183

and Archaea handle this challenge differently due to the dissimilar structures of their membrane 184

lipids (reviewed in detail by Oger and Cario (2013), Koga and Morii (2005), Koga (2012), and 185

Mansilla et al. (2004)). We will only focus on bacterial lipids here. For a review on archaeal lipids 186

see Oger and Cario (2013).

187

Bacterial polar membrane lipids consist mainly of straight-chain fatty acids that are bound 188

to the polar head group predominantly by ester linkages (Koga and Morii 2005). Bacteria respond 189

to various temperatures by altering the composition (length, degree of branching and degree of 190

unsaturation) of their fatty acid chains to maintain membrane fluidity (Mansilla et al. 2004; Zhang 191

and Rock 2008). The types of fatty acids bacteria are able to produce will therefore influence the 192

temperature range within which they can grow. For example, hyperthermophilic Thermotogae 193

have unusual membrane-spanning diabolic fatty acids in their membrane, which are thought to be 194

an adaptation to high temperature growth (Carballeira et al. 1997; Damsté et al. 2007). In 195

agreement with this hypothesis, these diabolic fatty acids are not found in the membranes of the 196

mesophilic Mesotoga prima (Nesbø et al. 2012). Moreover, M. prima (grown at 35°C) contained 197

branched, mono-unsaturated and saturated fatty acids, while K. olearia (grown at 55°C) contained 198

only saturated fatty acids (Nesbø et al. 2012). Fatty acid composition is also part of the immediate 199

cold-shock response with genes involved in production of, for instance, branched fatty-acids 200

(10)

10 being up-regulated in the thermophile Thermoanaerobacter tengcongensis when grown at sub- 201

optimal temperatures (Liu et al. 2014). Increase of branched fatty acids is a common response to 202

lower temperatures (Suutari and Laakso 1994), and in Listeria monocytogenes this is due to 203

temperature-dependent substrate selectivity of FabH, the enzyme responsible for the first 204

condensation reaction in fatty acid biosynthesis (Singh et al. 2009). Interestingly, in Bacillus a 205

transmembrane two-component response regulator, which controls the desaturase that introduces 206

double bonds in preexisting fatty acids, senses changes in membrane fluidity and not the actual 207

temperature changes (de Mendoza 2014).

208

In addition to the lipid structure of cell membranes, integral membrane proteins affect the 209

temperature tolerance of an organism (Thompkins et al. 2008). Therefore, while the lipid 210

composition of the membrane is crucial for its function, integral membrane proteins may also 211

play a significant role, particularly with respect to the temperature limit of an organism’s growth 212

range. For example, mutations of integral membrane proteins of the DedA family cause 213

temperature sensitivity and cell division defects in Escherichia coli (Thompkins et al. 2008).

214

Interestingly, proteins from the DedA family have been shown to be essential in at least two 215

bacterial species (E. coli and Borrelia burgdorferi), but their homologs are not detected in 216

several thermophilic and hyperthermophilic Thermotogae genomes (Doerrler et al. 2013). This 217

suggests that the function provided by DedA is either not needed by these organisms, or is being 218

provided by analogous integral membrane proteins, or that their DedA homologs are too 219

divergent to be detected by sequence similarity searches.

220 221

Nucleic acids: a challenge to keep the strands together 222

(11)

11 High temperatures denature double stranded DNA and secondary structures of RNA. This

223

presents a problem for thermophiles, and for hyperthermophiles in particular. These organisms 224

must maintain their chromosomes in an orderly state for both efficient packaging as well as 225

coordinated gene expression. Therefore, to survive the damaging effects of high temperature 226

thermophiles need to either continuously repair their damaged DNA or protect it from damage in 227

the first place. For example, the archaeon Pyrococcus abyssi has a highly efficient DNA repair 228

system that continuously repairs temperature-induced DNA damage (Jolivet et al. 2003). Very 229

high levels of homologous recombination are observed in hyperthermophilic Thermotoga spp.

230

where the ratio of nucleotide changes introduced by recombination relative to point mutation 231

(r/m) is in the range 24-100 for genomes originating from geographically distant sites (Nesbø et 232

al. 2006a; Nesbø et al. 2014). This in the upper range of values reported in a comparison of r/m 233

across a large sample of mostly mesophilic Bacteria and Archaea (0.02 – 64), where values 234

above 10 were interpreted as very high (Vos and Didelot 2009). The high level of recombination 235

may be explained by the need for DNA repair in thermophiles (Johnston et al. 2014). This 236

hypothesis is supported by observations of high levels of recombination and repair in other 237

hyperthermophilic microorganisms, such as Pyrococcus furiosus (DiRuggiero et al. 1997), 238

Sulfolobus islandicus (Whitaker et al. 2005), and Persephonella (Mino et al. 2013).

239

Protection of DNA is known to occur via multiple unrelated mechanisms. Primarily, 240

thermophiles safeguard their DNA with thermostable proteins analogous to eukaryotic histones.

241

For example, in the archaeon Thermococcus kodakaraensis HpkA and HpkB dramatically 242

increase the melting temperature of a given DNA sequence upon binding, with HpkB being able 243

to raise the melting temperature of poly(dA-dT) DNA by > 20°C (Higashibata et al. 1999), 244

suggesting that these proteins play a major role in the stabilization of Thermococcus 245

(12)

12 kodakaraensis chromosomes. In the bacterium T. maritima the histone-like protein HU stabilizes 246

and protects the DNA (Mukherjee et al. 2008).

247

Thermophiles can also use polyamine compounds to stabilize their DNA and RNA, as 248

well as many other cellular components. Multivalent polyamine compounds such as putrescine, 249

spermidine, and spermine, or their acetylated forms, compact histone-bound DNA in 250

Thermococcus kodakaraensis, stabilizing it at temperatures as high as 90°C (Higashibata et al.

251

2000). In Thermotoga species the polyamines caldopentamine and caldohexamine increase in 252

concentration with increased temperature, suggesting a role in thermal response and thermal 253

adaptation (Zellner and Kneifel 1993). Indeed caldopentamine and caldohexamine, as well as five 254

other long linear polyamines found in Thermus thermophilus, have been shown to stabilize 255

double-stranded DNA at high temperature, with a greater stabilizing effect by polyamines with a 256

larger number of amino nitrogen atoms (Terui et al. 2005).

257

Thirdly, unique RNA modifications can confer thermostability in thermophiles 258

(McCloskey et al. 2001). For example, modifications from adenosine to 2’-O-methyladenosine or 259

from guanosine to N2,2’-O-dimethylguanosine in the tRNAs are often growth temperature- 260

specific, even among closely related lineages (McCloskey et al. 2001).

261

Lastly, thermal adaptation may be achieved via reverse gyrase-mediated DNA 262

supercoiling. Reverse gyrase is a protein found almost exclusively in hyperthermophiles and, 263

importantly, it is a gene carried by all known hyperthermophiles (Brochier-Armanet and Forterre 264

2006; Forterre 2002; Lulchev and Klostermeier 2014). While deletion of the reverse gyrase gene 265

from Thermococcus kodakaraensis results in slower growth at high temperatures (90°C), it does 266

not abolish its growth, suggesting that this enzyme is not essential for hyperthermophilic growth 267

as was once thought (Atomi et al. 2004). However, since the T. kodakaraensis mutant lacking 268

(13)

13 reverse gyrase grew poorly at 90°C, and unlike the wild-type strain, could not grow above 90°C 269

(Atomi et al. 2004), this enzyme is still considered to be a critical adaptation for optimal growth 270

at high temperatures (Brochier-Armanet and Forterre 2006). Although reverse gyrase catalyzes 271

ATP-dependent positive supercoiling of DNA in vitro, its function in vivo remains unknown. The 272

increased heat protection provided by this enzyme may be linked to a role in the DNA damage 273

response, possibly through recruitment to lesions (Lulchev and Klostermeier 2014; Perugino et al.

274

2009). Interestingly, cultivated hyperthermophilic species from both the Thermotogae and the 275

Aquificae have acquired their reverse gyrase genes from Archaea by lateral gene transfer (LGT), 276

suggesting that hyperthermophily may have been acquired by Bacteria from Archaea (Brochier- 277

Armanet and Forterre 2006; Forterre et al. 2000).

278

While some of these adaptations for nucleic acid stabilization have only been found in 279

thermophiles (e.g., reverse gyrase (Forterre 2002), certain RNA modifications (McCloskey et al.

280

2001) and thermostable histones (Higashibata et al. 1999)), others are found in mesophiles as well.

281

For instance, the same polyamines found in Thermotoga are also found in mesophilic microalgae 282

(Nishibori et al. 2009). Hence, transition between thermophily and mesophily may only require a 283

re-purposing of certain cellular constituents, rather than removing or acquiring them.

284

In addition to cellular components interacting with nucleic acids for stabilization, the 285

composition of some nucleic acids appears adapted to the thermophilic lifestyle of the host 286

organism. The extra hydrogen bond in G:C nucleotide pairs was long thought to play a part in 287

optimal growth temperature. While genome-wide G+C content does not correlate with optimal 288

growth temperature (Galtier and Lobry 1997; Hurst and Merchant 2001; Zeldovich et al. 2007), 289

the G+C content of some structural RNA encoding genes does. For example, the G+C content of 290

secondary structures of rRNA and tRNA molecules, specifically in the stem structures, increases 291

(14)

14 with optimal growth temperature (Galtier and Lobry 1997; Kimura et al. 2013; Zhaxybayeva et al.

292

2009). As a result, the GC content variation of the 16S rRNA gene can be used as a proxy for 293

studying temperature adaptation within the Thermotogae. For example, the temperature optimum 294

for uncultured members of the phylum was predicted by establishing a correlation between the 295

16S rRNA gene distances and optimal growth temperature of 33 Thermotogae isolates (Dahle et 296

al. 2011). Additionally, inference of the ancestral states of the 16S rRNA gene that gave rise to 30 297

Thermotogae isolates allowed Green et al. (2013) to hypothesize that the thermotolerant 298

Thermotogae lineages are under directional selection and that transition from high to low optimal 299

growth temperature is easier to achieve.

300 301

Compatible solutes: the power of redundancy 302

Compatible solutes are organic compounds that are accumulated by cells under stressful 303

conditions such as osmotic stress and heat stress (Santos et al. 2011). These compounds, 304

particularly polyamines, are known to stabilize nucleic acids in thermophilic cells (see above).

305

Moreover, in the bacterium Calderobacterium hydrogenophilum polyamine compounds stabilize 306

the 70S initiation complex of ribosomes (Mikulik and Anderova 1994). Many temperature studies 307

in the Thermotogae have focused on the accumulation of these organic compounds and 308

polyamines and the elucidation of their biosynthetic pathways in T. maritima and the more 309

moderate thermophile Petrotoga miotherma (Jorge et al. 2007; Oshima et al. 2011; Rodionova et 310

al. 2013; Rodrigues et al. 2009; Zellner and Kneifel 1993). Several compatible solutes have so far 311

only been found in thermophiles including di-myo-inositol phosphate, mannosyl-di-myo-inositol 312

phosphate, mannosylglyceramide, and diglycerol phosphate (Borges et al. 2010; Gonçalves et al.

313

2012) and novel thermophilic solutes continue to be identified (Jorge et al. 2007; Rodrigues et al.

314

(15)

15 2009). However, while these compounds are thermophile-specific and may represent thermophile- 315

specific adaptations, they are not the only compatible solutes used to deal with heat stress. When 316

the ability to synthesize di-myo-inositol phosphate was removed from Thermococcus 317

kodakarensis by deleting a key synthesis gene, the growth of this archaeon was unaffected, and 318

aspartate accumulated as an alternative compatible solute (Borges et al. 2010). In the 319

Thermotogae multiple solutes accumulate under stress conditions (Jorge et al. 2007; Rodrigues et 320

al. 2009). This suggests that although the role compatible solutes play in thermal protection is not 321

fully understood, there is functional redundancy among the solutes.

322 323

Protein dynamics and turnover; assistance from chaperones and proteases 324

Chaperones are large protein complexes that assist the proper folding and re-folding of proteins.

325

The chaperonins represent an extensively studied subclass of chaperones with a stacked double-ring 326

structure (Large et al. 2009). Distribution of the chaperone families varies across Bacteria and 327

Archaea, and some chaperones are considered indispensable (Large et al. 2009). For example, some 328

chaperonins help fold new polypeptides, as well as re-fold and rescue proteins that have been 329

inactivated due to stress (Techtmann and Robb 2010). A major stressor that triggers chaperone- 330

mediated protein repair is heat shock, which has resulted in many chaperones being named heat 331

shock proteins (HSP) (Large et al. 2009). By preventing inactivation and aggregation of proteins at 332

high temperatures, this ubiquitous system is thought to be especially important in thermophiles, 333

which employ chaperones in both unstressed and heat-stressed states (Pysz et al. 2004). Thus, while 334

these proteins are part of high temperature response in mesophiles, their constitutive expression in 335

thermophiles may be part of their temperature adaptation. For example, the predicted chaperone 336

TM1083 in T. maritima is thought to stabilize the DNA gyrase enzyme at temperatures near optimal 337

(16)

16 growth (Canaves 2004). Moreover, the molecular chaperone trigger factor (TM0694) from T.

338

maritima strongly binds model proteins and decreases their folding rate, while these activities are 339

much weaker in the homologous trigger factor from the psychrophile Pseudoalteromonas 340

haloplanktis, which instead shows increased prolyl isomerization (Godin-Roulling et al. 2014).

341

However, it should be noted that chaperones, although always highly expressed in thermophiles, are 342

part of their high temperature response as well. For instance, examination of the T. maritima 343

proteome at four temperatures spanning its growth range revealed higher relative abundance of 344

chaperones at supra-optimal temperatures (Wang et al. 2012).

345

Proteases are also part of the heat shock response in mesophilic organisms (Richter et al.

346

2010). A key distinction between well-studied bacterial mesophiles and the hyperthermophile T.

347

maritima is the lack of regulation in T. maritima of most of its proteases in response to 348

temperature stress (Conners et al. 2006). This may be explained by an absence of major 349

regulators of the mesophilic proteolytic response (i.e., rpoH or ctsR homologs) in the T. maritima 350

genome (Conners et al. 2006; Pysz et al. 2004). Perhaps this bacterium gains a survival 351

advantage from constitutive expression of most proteases. A similarity search revealed an 352

absence of detectable rpoH and ctsR homologs in 38 Thermotogae, including the thermophilic K.

353

olearia and the mesophilic M. prima, suggesting that any regulation of protease expression in the 354

Thermotogae involves different genes than those used by other Bacteria and Archaea.

355

356

Thermal adaptation at the protein level 357

Although chaperones aid in proper folding and maintenance of proteins under high temperature 358

conditions, proteins from thermophilic organisms are themselves adapted to high temperature.

359

This adaptation is required to maintain activity at temperatures that would denature mesophilic 360

(17)

17 homologs and is found at all levels of protein structure, from primary through quaternary. Protein 361

thermostability is also not uniform across the proteome and depends on its functional role:

362

proteins either having catalytic activity or regulating other catalytic proteins appear to be under 363

greater selection to be temperature adapted than proteins involved in, for example, core 364

transcriptional or translational processes (Gu and Hilser 2009).

365

While there are many examples of specific thermostabilizing characteristics and 366

interactions at each of the four levels of globular protein structure (reviewed by Imanaka (2011) 367

and Li et al. (2005)), there is no universal property that confers thermostability. Rather, it is the 368

combination of factors at all levels of structure that grants high temperature activity in globular 369

proteins. Increased thermostability is often due to slight differences in sequence and structure, and 370

thermophilic and mesophilic counterparts are typically very similar proteins (Taylor and Vaisman 371

2010). Below we briefly overview known pathways to temperature adaptation in globular proteins.

372

Protein primary structure is the amino acid sequence of the polypeptide chain. Ultimately, 373

the properties and sequence of the amino acids determine the final higher level structures of the 374

protein. One characteristic associated with thermostable proteins is enrichment of amino acids that 375

contribute to a strong hydrophobic core. Larger aliphatic amino acids with more branches are 376

favored at positions that fill cavities, which may ultimately strengthen the protein through 377

increased hydrophobic interactions (Clark et al. 2004). Taylor and Vaisman (2010), however, 378

found that it is only a moderately good indicator of protein thermostability.

379

Comparisons of amino acid composition of thermophilic and mesophilic proteins have 380

revealed several trends at the global proteome level. The observed excess of charged (D,E,K,R) 381

versus polar (N,Q,S,T) amino acids in soluble proteins from hyperthermophiles, known as the 382

CvP bias (Cambillau and Claverie 2000; Gao and Wang 2012; Holder et al. 2013; Suhre and 383

(18)

18 Claverie 2003), may reflect larger importance of ionic interactions between charged amino acids 384

over hydrogen-bond interactions for retaining protein structure as temperature increases 385

(Cambillau and Claverie 2000). Additionally, a systematic evaluation of all possible subsets of 386

amino acids revealed that the total fraction of the amino acids IVYWREL in a proteome most 387

strongly correlates with optimal growth temperature (Zeldovich et al. 2007).

388

The CvP and IVYWREL biases have been explored thoroughly in the Thermotogae where 389

both indices show strong linear correlations with optimal growth temperature (Zhaxybayeva et al.

390

2009). Specifically, the distribution of CvP values was unimodal for each of the Thermotogae 391

proteomes, arguing against the hypothesis that thermophily is a recently acquired trait of the 392

Thermotogae. Moreover, calculation of CvP values from estimated ancestral Thermotogae 393

sequences suggested that the ancestral Thermotogae proteome belonged to organisms with an 394

optimal growth temperature of ≈84.5°C, higher than that of any characterized extant Thermotogae 395

bacterium (Zhaxybayeva et al. 2009). While the average CvP value for most of the thermophilic 396

Thermotogae lineages was above 10.62, the mesophilic M. prima proteome has an average CvP 397

value of 8.96 (Zhaxybayeva et al. 2012). Also this genome has a unimodal CvP distribution, 398

suggesting it has maintained a mesophilic lifestyle for a long time. An exception to the trend is 399

observed in the P. lettingae genome, which has an average CvP value of 8.42 (Zhaxybayeva et al.

400

2009), but an optimal growth temperature of 65°C. However, P. lettingae-like 16S rRNA genes 401

and genomic DNA have been recovered from environments with temperatures < 65°C (e.g., 40- 402

50°C, (Nesbø et al. 2010; Nobu et al. 2014)), suggesting that these bacteria often live at 403

temperatures below the optimal growth temperature of the cultivated isolate.

404

Protein secondary structure describes the local folding of polypeptide sequences. This 405

includes regular structures like α-helices and β-sheets, or irregular structures like β-turns, coils 406

(19)

19 and loops. These are formed primarily by hydrogen bond interactions between the backbone and 407

side chain elements of the amino acids. In addition to having secondary structures that facilitate 408

tighter packing and rigidity at the tertiary level, thermophilic proteins tend to have secondary 409

structures that are more stabilized than their mesophilic counterparts (Facchiano et al. 1998; Koga 410

et al. 2008; Prakash and Jaiswal 2010). For example, thermostable proteins have been reported to 411

have a larger fraction of their amino acid residues arranged in α-helices than mesophilic proteins 412

do (Prakash and Jaiswal 2010).

413

Protein tertiary structure is the arrangement of a folded polypeptide chain in three- 414

dimensional space. This is achieved by disulfide bridges, electrostatic interactions within the 415

polypeptide chain, and hydrophobic interactions and hydrogen bonding within the chain as well as 416

between the peptides and solvent. Thermophilic proteins tend to have conformations that are more 417

rigid and more tightly packed, with reduced entropy of unfolding and conformational strain 418

compared to their mesophilic counterparts (Li et al. 2005). The strongest contributors to 419

thermostability are increased ion pairs on the protein surface combined with a more strongly 420

hydrophobic interior (Taylor and Vaisman 2010). In agreement with this, additional salt bridges 421

on the surface of the enzyme diguanylate cyclase from T. maritima accounted for its greater 422

thermostability compared to the same enzyme found in the mesophiles Pseudomonas aeruginosa, 423

Marinobacter aquaeolei and Geobacter sulfurreducens (Deepthi et al. 2014). Additionally, the 424

glutamate dehydrogenase enzymes of the hyperthermophilic bacterium T. maritima and 425

hyperthermophilic archaeon P. furiosus have smaller hydrophobic accessible surface area (ASA) 426

and greater charged ASA than the glutamate dehydrogenase from the mesophilic bacterium 427

Clostridium symbiosum (Knapp et al. 1997). Since few other structural differences were found 428

(20)

20 between the thermophilic and mesophilic enzymes, this tighter packing is thought to contribute to 429

the thermal stability of the proteins.

430

Protein quaternary structure is the arrangement of multiple folded polypeptide chains into 431

a multimeric complex. In globular proteins this level of structure is formed and maintained by 432

many of the same forces that contribute to the tertiary structure of a protein, but between 433

polypeptide chains rather than within them. These forces include disulfide bridges, electrostatic 434

interactions, hydrophobic interactions and hydrogen bonding. In thermostable proteins, greater 435

numbers of these interactions, or stronger interactions over weaker ones, are favored (Li et al.

436

2005).

437

One additional way of achieving greater protein stability is to increase the number of 438

subunits. For example, the malate dehydrogenase (MDH) enzyme, which is usually a dimer in 439

mesophiles, is a tetramer in the thermophilic bacterium Chloroflexus aurantiacus (Bjørk et al.

440

2003). The additional dimer-dimer interface of the tetrameric MDH is hypothesized to provide 441

thermal stability due to the higher number of inter-polypeptide interactions compared to the 442

mesophilic dimers. To test this hypothesis, Bjørk et al. (2003) introduced a disulfide bridge that 443

would strengthen dimer-dimer interaction further, and found that the new enzyme had a melting 444

temperature 15°C higher than the wild-type enzyme. In addition, removing excess negative charge 445

at the dimer-dimer interface by replacing a glutamate residue with either glutamine or lysine 446

resulted in an increase of apparent melting temperature by ~ 24°C (Bjørk et al. 2004).

447 448

Tolerating new temperatures: Is it possible to modify just a few proteins?

449

As discussed above, adaptation to a high optimal growth temperature is achieved differently by 450

Bacteria and Archaea, by one species than another, and even by one protein than another within 451

(21)

21 the same organism. Given that all of these factors combine in unique ways to permit growth 452

within a specific temperature range, how could a shift in permissive temperature range be 453

accomplished? While some of these strategies are universal to thermophiles and mesophiles, such 454

as utilization of chaperones and compatible solutes, others, like shifting of membrane properties, 455

would have to be radically altered to accommodate large changes in temperature range.

456

Changing a few key proteins may have global stabilizing effects on the whole cell. For 457

instance, some of the proteins whose stability appears most affected by thermal adaptation are 458

involved in production of compatible solutes that stabilize other proteins (Gu and Hilser 2009).

459

Such changes would reduce the need to modify the stability of all components of the proteome. It 460

may also be possible to lower the maximal growth temperature of an organism through changes to 461

a single protein (Endo et al. 2006). By replacing the chromosomal copy of groEL chaperonin in 462

Bacillus subtilis 168 (growth range from 11 to 52°C) with a psychrophilic groEL from 463

Pseudoalteromonas sp. PS1M3 (growth range from 4 to 30°C), Endo and colleagues noted a 2°C 464

reduction in the maximal growth temperature of the newly constructed B. subtilis strain. Similarly, 465

the heterologous expression of a small heat shock protein from Caenorhabditis elegans, enabled E.

466

coli cells to grow at temperatures up to 50°C (and survive heat shock at 58°C for 1/2h) extending 467

its growth range by 3.5°C (Ezemaduka et al. 2014). While these changes do not constitute true 468

shifts in growth temperature range or changes to optimal growth temperature, these studies 469

suggest that changes to a single key protein (involved both in temperature adaptation and 470

response) could extend or narrow the temperature range at which an organism is able to grow by a 471

few degrees. Accumulation of several such mutations could eventually lead to a more substantial 472

shift in growth range. Some of these mutations may be advantageous at lower temperatures, while 473

others may be loss-of-function mutations, where abilities to function at higher temperatures are 474

(22)

22 lost for proteins in individuals living in an environment with temperatures at the lower end of 475

their original growth range. Under the latter scenario, change in the growth temperature range 476

might not be a result of selection, but rather a product of random genetic drift or genetic 477

hitchhiking with another, unrelated trait selected for in the new environment.

478 479

Role of Lateral Gene Transfer in Temperature Adaptation: Acquisition of Already 480

‘Adapted’ Genes 481

Lateral gene transfer (LGT) is a major force in prokaryotic evolution, allowing rapid adaptation to 482

changes in the environment by acquiring clusters of genes or single genes that confer a selective 483

advantage (Boucher et al. 2003; Zhaxybayeva and Doolittle 2011) and LGT has been implicated 484

in adaptation to extreme environments including high temperatures (see for example Omelchenko 485

et al. (2005)). Genes encoding proteins linked to adaptation to higher or lower growth 486

temperatures have been laterally exchanged (reviewed in Boucher et al. 2003). Reverse gyrase is 487

a classic example of lateral transfer of a single gene that is thought to have been crucial for 488

evolutionary adaptation to high temperatures by hyperthermophilic Bacteria (Brochier-Armanet 489

and Forterre 2006; Forterre 2002). Phylogenetic analyses suggest two ancient acquisitions of this 490

gene by bacterial lineages from Archaea, followed by secondary transfer events among Bacteria 491

(Brochier-Armanet and Forterre 2006).

492

Similarly, the compatible solute di-myo-inositol phosphate is thought to be important for 493

heat tolerance in thermophiles and hyperthermophiles (Borges et al. 2010). Two key genes 494

involved in the synthesis of this compound (inositol-1-phosphate cytidylyltransferase and di-myo- 495

inositol phosphate phosphate synthase) are suggested to have been laterally transferred from an 496

archaeal lineage to hyperthermophilic marine Thermotoga species, while in other lineages the two 497

(23)

23 genes are predicted to have fused before being exchanged among several bacterial and archaeal 498

lineages (Gonçalves et al. 2012).

499

Reverse gyrase and the myo-inositol pathway genes are just two examples of a large 500

number of genes transferred into the Thermotogae. When the genome of T. maritima MSB8 was 501

first sequenced (Nelson et al. 1999), 24% of its open reading frames (ORFs) showed greatest 502

similarity to sequences from Archaea, suggesting that these genes have been acquired from these 503

distantly related organisms that inhabit the same environment. Comparative genomic analyses of 504

additional Thermotogae genomes have confirmed an influx of genes from Archaea (albeit the total 505

number dropped to 10-11% of the ORFs, due to increased number of bacterial homologs in 506

GenBank) and an even larger fraction of Firmicutes genes in these genomes (Mongodin et al.

507

2005; Nesbø et al. 2009; Zhaxybayeva et al. 2009; Zhaxybayeva et al. 2012). Phylogenetic 508

analysis of all the ORFs in the M. prima genome suggests this lineage has undergone extensive 509

gene exchange with diverse mesophilic lineages, and that LGT has aided its transition from a 510

thermophilic to a mesophilic lifestyle (Zhaxybayeva et al. 2012). Thus, as a major force that has 511

shaped the genomes of the Thermotogae, LGT may have also been important for the acquisition 512

and development of the temperature ranges of the various Thermotogae lineages. Most of the 513

acquired genes in Thermotogae (including M. prima) are involved in carbohydrate metabolism 514

(Mongodin et al. 2005; Nesbø et al. 2009; Zhaxybayeva et al. 2009; Zhaxybayeva et al. 2012).

515

However, M. prima has additionally acquired genes involved in signal transduction mechanisms, 516

secondary metabolite biosynthesis, and amino acid transport and metabolism (Zhaxybayeva et al.

517

2012), suggesting the potential importance of genes from these functional categories for life at 518

lower temperatures.

519 520

(24)

24 Transition to mesophily in Kosmotoga and Mesotoga

521

The discovery of the mesophilic Thermotogae lineage (Mesotoga) raised the possibility that 522

(hyper)thermophily was not ancestral to the phylum. However, as discussed above, the amino acid 523

composition (CvP bias and IVYWREL amino acids frequency) of the reconstructed ancestral 524

Thermotogae proteome suggests that the ancestral Thermotogae was a hyperthermophile 525

(Zhaxybayeva et al. 2009), and that the transition to mesophily in the Thermotogae phylum is 526

secondary. Moreover, ancestral sequence reconstruction of myo-inositol-3-phosphate synthase 527

enzymes in the Thermotoga genus also suggests that the ancestor of this hyperthermophilic 528

lineage grew optimally at temperatures higher than those of extant species (Butzin et al. 2013).

529

The G+C content of ribosomal RNA, which correlates with optimal growth temperature, also 530

suggests that the reconstructed 16S rRNA of the ancestor of all Thermotogae belonged to a 531

thermophile (Green et al. 2013; Zhaxybayeva et al. 2009).

532

So far, the genus Mesotoga is the only strictly mesophilic Thermotogae, with optimal 533

growth occurring between 37 and 45°C (Ben Hania et al. 2013; Nesbø et al. 2012). Initially 534

Mesotoga spp. were only detected using molecular tools such as community 16S rRNA PCR and 535

metagenome analyses (Nesbø et al. 2006b). Mesotoga prima was the first described isolate of the 536

genus (Nesbø et al. 2012), which now includes another validated species, Mesotoga infera, (Ben 537

Hania et al. 2013), one yet to be validated, Mesotoga sp. PhosAc3 (Ben Hania et al. 2011), and 538

several isolates with ongoing genome sequencing projects (Benson et al. 2014; Reddy et al.

539

2014). The 2.97 Mb genome of M. prima is considerably larger than any previously sequenced 540

Thermotogae genome, which range between 1.86 and 2.30 Mb (Zhaxybayeva et al. 2012). This 541

larger size is due to both higher numbers of protein-coding genes and larger intergenic regions. A 542

unimodal distribution of CvP values of M. prima's proteome, with a mean value in the 543

(25)

25 mesophilic range, indicate that native M. prima proteins have also changed in response to its 544

evolved mesophilic lifestyle (Zhaxybayeva et al. 2012).

545

Analysis of additional Thermotogae shows that the variation in size may be related to 546

optimal growth temperature: thermophiles have more streamlined genomes, with little intergenic 547

space and a higher number of genes per transcription unit, while mesophiles have larger 548

intergenic spaces and higher gene redundancy (Latif et al. 2013; Zhaxybayeva et al. 2012). This 549

finding holds true for lineages outside of the Thermotogae, as examination of 1155 prokaryotes 550

demonstrates (Sabath et al. 2013). However, the observed correlation in Thermotogae needs to 551

be untangled from effects of phylogenetic history (Zhaxybayeva et al. 2012).

552

The closest relative of the Mesotoga lineage is the thermophilic lineage Kosmotoga (Fig.

553

3). Members of this genus have been found in hydrothermal sediments (L'Haridon et al. 2014;

554

Nunoura et al. 2010) and oil production fluids (DiPippo et al. 2009; Feng et al. 2010). Like other 555

thermophilic Thermotogae, the Kosmotoga are anaerobic chemoorganotrophs able to ferment 556

carbohydrates and peptides (Nunoura et al. 2010) and to produce molecular hydrogen (DiPippo 557

et al. 2009; Feng et al. 2010). The first isolated bacterium of this genus was Kosmotoga olearia 558

(DiPippo et al. 2009). K. olearia grows optimally at 65°C and has a reported growth range of 20- 559

80°C (DiPippo et al. 2009). Not only is this bacterium capable of growing at an unusually low 560

temperature for a thermophile, but to our knowledge it represents the widest reported bacterial 561

temperature growth range to date.

562

The ability of Kosmotoga to grow at such an extraordinary gamut of temperatures is 563

intriguing for two reasons. First, it must maintain protein activity and membrane integrity. Every 564

living organism has adapted to do this at a certain temperature range, but how these requirements 565

can be maintained over a 60°C range is unknown. What evolutionary mechanisms would maintain 566

(26)

26 a 60°C growth interval in Kosmotoga? Perhaps this lineage continues to experience environments 567

with more variable temperatures or, alternatively, the wide growth range may be a result of 568

selection on another trait. Second, as discussed above, this ability of tolerating a wide range of 569

temperature conditions, may have facilitated the transition of Mesotoga from thermophily to 570

mesophily, because the capacity to grow at lower temperatures presumably already existed in 571

Mesotoga's ancestors.

572

As a result Kosmotoga and Mesotoga offer a unique model system for studying both 573

immediate temperature responses and long-term temperature adaptation. Specifically, K. olearia’s 574

exceptionally wide growth range allows examination of temperature responses under both 575

mesothermic and thermic conditions in the same cell-line. For example, analysis of K. olearia’s 576

transcriptome at different growth temperatures promises to shed light into the role of specific 577

processes, functions, genes or proteins in thermoadaptation. Since K. olearia’s closest relative is a 578

mesophile with a narrower growth range, comparative genomic, transcriptomic and proteomic 579

analyses promise to reveal how Kosmotoga's temperature responses may eventually lead to 580

temperature adaptation. If we assume that Mesotoga and Kosmotoga’s common ancestor was a 581

thermophile, possibly with a wide growth range, then the Mesotoga lineage lost its ability to grow 582

at high temperatures, while Kosmotoga has either kept or expanded its growth range. For 583

Mesotoga we have speculated that reduction of its growth temperature range may have happened 584

as the lineage got 'trapped' in an oil reservoir that cooled down (Nesbø et al. 2006b; Zhaxybayeva 585

et al. 2012) and therefore may be a result of loss-of-function mutations and genetic drift.

586

The existence of several additional Thermotogae lineages likely thriving in mesothermic 587

environments (Nesbø et al. 2010) opens opportunities to study the evolutionary processes in 588

lineages that have adapted to lower temperatures independently. These novel lineages can be 589

(27)

27 accessed through metagenomic studies or through further cultivation efforts. Taken together, 590

future genomic, transcriptomic and proteomic studies of temperature responses and adaptations 591

in Kosmotoga, Mesotoga, and other Thermotogae will help decipher how shifts in temperature 592

range and optimum are accomplished.

593 594

Acknowledgements 595

This work is supported by an NSERC Alexander Graham Bell Canada Graduate Scholarship 596

CGS-M to S.M.J.P., by a Norwegian Research Council of Norway award (project no.

597

180444/V40) to C.L.N., and by a Simons Investigator award from the Simons Foundation, 598

Dartmouth’s Walter and Constance Burke Research Initiation Award and Dean of Faculty start- 599

up funds to O.Z.

600 601

References 602

603

Achenbach-Richter, L., Gupta, R., Stetter, K.O., and Woese, C.R. 1987. Were the original 604

eubacteria thermophiles? Syst. Appl. Microbiol. 9(1-2): 34-39.

605

Adl, S.M., Simpson, A.G.B., Lane, C.E., Lukeš Julius, Bass, D., Bowser, S.S. et al. 2012. The 606

revised classification of eukaryotes. J. Eukaryot. Microbiol. 59(5): 429-493.

607

Akanuma, S., Nakajima, Y., Yokobori, S., Kimura, M., Nemoto, N., Mase, T., et al. 2013.

608

Experimental evidence for the thermophilicity of ancestral life. Proc. Natl. Acad. Sci. U. S. A.

609

110(27): 11067-11072.

610

Arti, D., Park, J., Jung, T.Y., Song, H., Jang, M., Han, N.S. et al. 2012. Structural analysis of α- 611

L-Arabinofuranosidase from Thermotoga maritima reveals characteristics for thermostability and 612

substrate specificity. J.Microbiol.Biotechnol. 22(12): 1724-1730.

613

Atomi, H., Matsumi, R., and Imanaka, T. 2004. Reverse gyrase is not a prerequisite for 614

hyperthermophilic life. J. Bacteriol. 186(14): 4829-4833.

615

(28)

28 Ben Hania, W., Ghodbane, R., Postec, A., Brochier-Armanet, C., Hamdi, M., Fardeau, M. et al.

616

2011. Cultivation of the first mesophilic representative (mesotoga) within the order 617

Thermotogales. Syst. Appl. Microbiol. 34(8): 581-585.

618

Ben Hania, W., Postec, A., Aüllo, T., Ranchou-Peyruse, A., Erauso, G., Brochier-Armanet, et al.

619

2013. Mesotoga infera sp. nov., a mesophilic member of the order Thermotogales, isolated from 620

an underground gas storage aquifer. Int. J. Syst. Evol. Microbiol. 63: 3003-3008.

621

Benson, D.A., Clark, K., Karsch-Mizrachi, I., Lipman, D.J., Ostell, J., and Sayers, E.W. 2014.

622

GenBank. Nucleic Acids Res. 43: D30-D35.

623

Bhandari, V., and Gupta, R.S. 2014. Molecular signatures for the phylum (class) Thermotogae 624

and a proposal for its division into three orders (Thermotogales, Kosmotogales ord. nov. and 625

Petrotogales ord. nov.) containing four families (Thermotogaceae, Fervidobacteriaceae fam.

626

nov., Kosmotogaceae fam. nov. and Petrotogaceae fam. nov.) and a new 627

genus Pseudothermotoga gen. nov. with five new combinations. Antonie van Leeuwenhoek 628

105(1): 143-168.

629

Bjørk, A., Dalhus, B., Mantzilas, D., Sirevåg, R., and Eijsink, V.G.H. 2004. Large improvement 630

in the thermal stability of a tetrameric malate dehydrogenase by single point mutations at the 631

dimer-dimer interface. J. Mol. Biol. 341(5): 1215-1226.

632

Bjørk, A., Dalhus, B., Mantzilas, D., Eijsink, V.G.H., and Sirevåg, R. 2003. Stabilization of a 633

tetrameric malate dehydrogenase by introduction of a disulfide bridge at the dimer-dimer 634

interface. J. Mol. Biol. 334(4): 811-821.

635

Borges, N., Matsumi, R., Imanaka, T., Atomi, H., and Santos, H. 2010. Thermococcus 636

kodakarensis mutants deficient in di-myo-inositol phosphate use aspartate to cope with heat 637

stress. J. Bacteriol. 192(1): 191-197.

638

Boucher, N., and Noll, K.M. 2011. Ligands of thermophilic ABC transporters encoded in a 639

newly sequenced genomic region of Thermotoga maritima MSB8 screened by differential 640

scanning fluorimetry. Appl. Environ. Microbiol. 77(18): 6395-6399.

641

Boucher, Y., Douady, C.J., Papke, R.T., Walsh, D.A., Boudreau, M.E.R., Nesbø, C.L., et al.

642

2003. Lateral gene transfer and the origins of prokaryotaic groups. Annu. Rev. Genet. 37: 283- 643

328.

644

Boussau, B., Blanquart, S., Necsulea, A., Lartillot, N., and Gouy, M. 2008. Parallel adaptations 645

to high temperatures in the Archaean eon. Nature 456(7224): 942-945.

646

Brochier-Armanet, C., and Forterre, P. 2006. Widespread distribution of archaeal reverse gyrase 647

in thermophilic bacteria suggests a complex history of vertical inheritance and lateral gene 648

transfers. Archaea 2(2): 83-93.

649

(29)

29 Butzin, N.C., Lapierre, P., Green, A.G., Swithers, K.S., Gogarten, J.P., and Noll, K.M. 2013.

650

Reconstructed Ancestral Myo-Inositol-3-Phosphate Synthases Indicate That Ancestors of the 651

Thermococcales and Thermotoga Species Were More Thermophilic than Their Descendants.

652

PloS one 8(12): e84300.

653

Cambillau, C., and Claverie, J. 2000. Structural and genomic correlates of hyperthermostability.

654

The Journal of biological chemistry 275(42): 32383-32386.

655

Canaves, J.M. 2004. Predicted role for the Archease protein family based on structural and 656

sequence analysis of TM1083 and MTH1598, two proteins structurally characterized through 657

structural genomics efforts. Proteins 56(1): 19-27.

658

Cappelletti, M., Zannoni, D., Postec, A., and Ollivier, B. 2014. Members of the order 659

Thermotogales: From microbiology to hydrogen production. In Microbial BioEnergy: Hydrogen 660

Production. Edited by D. Zannoni and R. De Philippis. Springer Netherlands, Dordrecht. pp. 197- 661

224.

662

Carballeira, N.M., Reyes, M., Sostre, A., Huang, H., Verhagen, M.F., and Adams, M.W. 1997.

663

Unusual fatty acid compositions of the hyperthermophilic archaeon Pyrococcus furiosus and the 664

bacterium Thermotoga maritima. J. Bacteriol. 179(8): 2766-2768.

665

Chang, E.L. 1994. Unusual thermal stability of liposomes made from bipolar tetraether lipids.

666

Biochem. Biophys. Res. Commun. 202(2): 673-679.

667

Chhabra, S.R., Shockley, K.R., Ward, D.E., and Kelly, R.M. 2002. Regulation of endo-acting 668

glycosyl hydrolases in the hyperthermophilic bacterium Thermotoga maritima grown on glucan- 669

and mannan-based polysaccharides. Appl. Environ. Microbiol. 68: 545-554.

670

Chhabra, S.R., Shockley, K.R., Conners, S.B., Scott, K.L., Wolfinger, R.D., and Kelly, R.M.

671

2003. Carbohydrate-induced differential gene expression patterns in the hyperthermophilic 672

bacterium Thermotoga maritima. J. Biol. Chem. 278(9): 7540-7552.

673

Clark, A.T., McCrary, B.S., Edmondson, S.P., and Shriver, J.W. 2004. Thermodynamics of core 674

hydrophobicity and packing in the hyperthermophile proteins Sac7d and Sso7d. Biochemistry 675

43: 2840-2853.

676

Cobucci-Ponzano, B., Zorzetti, C., Strazzulli, A., Carillo, S., Bedini, E., Corsaro, M.M. et al.

677

2011. A novel α-D-galactosynthase from Thermotoga maritima converts β-D-galactopyranosyl 678

azide to α-galacto-oligosaccharides. Glycobiology 21(4): 448-456.

679

Comfort, D.A., Bobrov, K.S., Ivanen, D.R., Shabalin, K.A., Harris, J.M., Kulminskaya, A.A. et 680

al. 2007. Biochemical analysis of Thermotoga maritima GH36 α-galactosidase (TmGalA) 681

confirms the mechanistic commonality of clan GH-D glycoside hydrolases. Biochemistry 682

46(11): 3319-3330.

683

(30)

30 Conners, S.B., Mongodin, E.F., Johnson, M.R., Montero, C.I., Nelson, K.E., and Kelly, R.M.

684

2006. Microbial biochemistry, physiology, and biotechnology of hyperthermophilic Thermotoga 685

species. FEMS Microbiol. Rev. 30(6): 872-905.

686

Cuneo, M.J., Beese, L.S., and Hellinga, H.W. 2009. Structural analysis of semi-specific 687

oligosaccharide recognition by a cellulose-binding protein of Thermotoga maritima reveals 688

adaptations for functional diversification of the oligopeptide periplasmic binding protein fold. J.

689

Biol. Chem. 284(48): 33217-33223.

690

Dagan, T., Roettger, M., Bryant, D., and Martin, W. 2010. Genome networks root the tree of life 691

between prokaryotic domains. Genome Biology and Evolution 2: 379-392.

692

Dahle, H., Hannisdal, B., Steinsbu, B.O., Ommedal, H., Einen, J., Jensen, et al. 2011. Evolution 693

of temperature optimum in Thermotogaceae and the prediction of trait values of uncultured 694

organisms. Extremophiles 15(4): 509-516.

695

Damsté, J.S.S., Rijpstra, W.I.C., Hopmans, E.C., Schouten, S., Balk, M., and Stams, A.J.M.

696

2007. Structural characterization of diabolic acid-based tetraester, tetraether and mixed 697

ether/ester, membrane-spanning lipids of bacteria from the order Thermotogales. Arch.

698

Microbiol. 188(6): 629-641.

699

de Mendoza, D. 2014. Temperature sensing by membranes. Annu. Rev. Microbiol. 68: 101-116.

700

Deepthi, A., Liew, C.W., Liang, Z., Swaminathan, K., and Lescar, J. 2014. Structure of a 701

diguanylate cyclase from Thermotoga maritima: Insights into activation, feedback inhibition and 702

thermostability. PloS one 9(10): e110912.

703

DiDonato, M., Deacon, A.M., Klock, H.E., McMullan, D., and Lesley, S.A. 2004. A scaleable 704

and integrated crystallization pipeline applied to mining the Thermotoga maritima proteome. J.

705

Struct. Funct. Genomics 5: 133-146.

706

DiPippo, J.L., Nesbø, C.L., Dahle, H., Doolittle, W.F., Birkland, N., and Noll, K.M. 2009.

707

Kosmotoga olearia gen. nov., sp. nov., a thermophilic, anaerobic heterotroph isolated from an oil 708

production fluid. Int. J. Syst. Evol. Microbiol. 59: 2991-3000.

709

D'Ippolito, G., Dipasquale, L., and Fontana, A. 2014. Recycling of carbon dioxide and acetate as 710

lactic acid by the hydrogen-producing bacterium Thermotoga neapolitana. ChemSusChem 7:

711

2678-2683.

712

DiRuggiero, J., Santangelo, N., Nackerdien, Z., Ravel, J., and Robb, F.T. 1997. Repair of 713

extensive ionizing-radiation DNA damage at 95 degrees C in the hyperthermophilic archaeon 714

Pyrococcus furiosus. J. Bacteriol. 179(14): 4643-4645.

715

Doerrler, W.T., Sikdar, R., Kumar, S., and Boughner, L.A. 2013. New functions for the ancient 716

DedA membrane protein family. J. Bacteriol. 195(1): 3-11.

717

(31)

31 Endo, A., Sasaki, M., Maruyama, A., and Kurusu, Y. 2006. Temperature adaptation of Bacillus 718

subtilis by chromosomal groEL replacement. Biosci. Biotechnol. Biochem. 70(10): 2357-2362.

719

Ezemaduka, A.N., Yu, J., Shi, X., Zhang, K., Yin, C., Fu, X. et al. 2014. A small heat shock 720

protein enables Escherichia coli to grow at a lethal temperature of 50°C conceivably by 721

maintaining cell envelope integrity. J. Bacteriol. 196(11): 2004-2011.

722

Facchiano, A., Colonna, G., and Ragone, R. 1998. Helix-stabilizing factors and stabilization of 723

thermophilic proteins: an X-ray based study. Protein Eng. 11(9): 753-760.

724

Feng, Y., Cheng, L., Zhang, X., Li, X., Deng, Y., and Zhang, H. 2010. Thermococcoides 725

shengliensis gen. nov., sp. nov., a new member of the order Thermotogales isolated from oil- 726

production fluid. Int. J. Syst. Evol. Microbiol. 60: 932-937.

727

Forterre, P., Bouthier De La Tour C, Philippe, H., and Duguet, M. 2000. Reverse gyrase from 728

hyperthermophiles: probable transfer of a thermoadaptation trait from Archaea to Bacteria.

729

Trends in genetics : TIG 16(4): 152-154.

730

Forterre, P. 2002. A hot story from comparative genomics: reverse gyrase is the only 731

hyperthermophile-specific protein. Trends in genetics : TIG 18(5): 236-237.

732

Frock, A.D., Gray, S.R., and Kelly, R.M. 2012. Hyperthermophilic Thermotoga species differ 733

with respect to specific carbohydrate transporters and glycoside hydrolases. Appl. Environ.

734

Microbiol. 78(6): 1978-1986.

735

Galtier, N., and Lobry, J.R. 1997. Relationships between genomic G+C content, RNA secondary 736

structures, and optimal growth temperature in prokaryotes. J. Mol. Evol. 44(6): 632-636.

737

Gao, J., and Wang, W. 2012. Analysis of structural requirements for thermo-adaptation from 738

orthologs in microbial genomes. Annals of Microbiology 62(4): 1635-1641.

739

Ghimire-Rijal, S., Lu, X., Myles, D.A., and Cuneo, M.J. 2014. Duplication of genes in an ATP- 740

binding cassette transport system increases dynamic range while maintaining ligand specificity.

741

J. Biol. Chem. 289(43): 30090-30100.

742

Godin-Roulling, A., Schmidpeter, P.A.M., Schmid, F.X., and Feller, G. 2014. Functional 743

adaptations of the bacterial chaperone trigger factor to extreme environmental temperatures.

744

Environ. Microbiol. in press.

745

Gogarten, J.P., Kibak, H., Dittrich, P., Taiz, L., Bowman, E.J., Bowman, et al. 1989. Evolution 746

of the vacuolar H+-ATPase: implications for the origin of eukaryotes. Proc. Natl. Acad. Sci. U.

747

S. A. 86(September): 6661-6665.

748

Gonçalves, L.G., Borges, N., Serra, F., Fernandes, P.L., Dopazo, H., and Santos, H. 2012.

749

Evolution of the biosynthesis of di-myo-inositol phosphate, a marker of adaptation to hot marine 750

environments. Environ. Microbiol. 14(3): 691-701.

751

Referanser

RELATERTE DOKUMENTER

Due to the thermoresponsive nature and lower critical solution temperature (LCST) it is interesting to first examine how the turbidity changes with temperature for these

3.1 Evolution of costs of defence 3.1.1 Measurement unit 3.1.2 Base price index 3.2 Operating cost growth and investment cost escalation 3.3 Intra- and intergenerational operating

Furthermore, we have identified the transporters responsible for GABA and tau- rine uptake in the liver by using isolated rat hepatocytes and by quantifying the levels of mRNAs

Inoperabilities ( q k ) for different Norwegian industry sectors that are caused by a notional 10% demand reduction for the sectors, together with cascading effects to other

Fig 12 Error in range estimate as function of global error in sound speed Red solid curve: 10 km range 40 degrees off broadside Blue dotted line: 10 km range 10 degrees off

We have applied the climate feedback re- sponses analysis method (CFRAM) to decompose the surface temperature changes into partial temperature contributions of individual

Due to the thermoresponsive nature and lower critical solution temperature (LCST) it is interesting to first examine how the turbidity changes with temperature for these samples.

In order to determine the extent to which demographics account for changes and dif- ferences in growth, both through changes in factor supplies and through changes in TFP,